You are on page 1of 11

JTTEE5 20:486–496

DOI: 10.1007/s11666-010-9546-4
1059-9630/$19.00  ASM International
Peer Reviewed

Mechanical Property Mapping of Cold


Sprayed Ti Splats and Coatings
Dina Goldbaum, Richard R. Chromik, Stephen Yue, Eric Irissou, and Jean-Gabriel Legoux

(Submitted April 15, 2010; in revised form August 1, 2010)

Profile nanoindentation and nanoindentation mapping were used to investigate the mechanical prop-
erties of commercially pure cold spray Ti splats and coatings deposited at increasing deposition veloci-
ties. Three regions in the cold spray Ti splats have been indentified: the impact region, the jetting region,
and the upper splat region. The mechanical properties measured in these regions were tied to the cold
spray deposition process with help of optical and scanning electron microscopes. The jetting region was
observed to contribute to a metallurgical bonding of cold spray splats to cold spray splats and was
measured to have low hardness in comparison to the splat impact site and similar to the hardness in the
upper splat region. No increase in the profile coatings hardness with increase in the particle in-flight
velocity and coating thickness was found. A correlation between the mechanical properties and the
presently known deposition temperature, stress and dislocation density models was made.

Keywords cold spray, hardness mapping, nanoindentation,


Zahiri et al. (Ref 14) as well as Vlcek et al. (Ref 19)
titanium reported a metallurgical material bonding that was
observed in the jetting region in cold sprayed Ti. Also,
Zahiri et al. (Ref 14) believed that high-temperature
fluctuations induced material softening and grain recrys-
tallization with preferential grain orientation in Ti cold
1. Introduction spray splats. Using transmission electron microscopy, Kim,
et al. showed both regions of high dislocation density at
Cold spray is an emerging coating deposition process the impact site of the splat with the substrate (Ref 23) and
(Ref 1) that is carried out at relatively low-temperature regions of recrystallization (Ref 24). Thus, a complex and
condition (Ref 2, 3). The deposition is achieved by powder varied microstructure is observed in cold sprayed Ti that
acceleration to supersonic velocities. Upon the particle may also result in a variation in local mechanical properties.
impact, the material undergoes plastic deformation which In this study, the hardness mapping of cold sprayed Ti
initially results in a conformal material adhesion (Ref 3). splats and coatings was used to investigate the presence or
At high enough deposition velocities, the particles lack of thermal softening, which could be associated with
undergo adiabatic shear instability where a localized material melting in the jetting region. The work carried
fluctuation in the pressure and strain rate leads to a out is based on a nanoindentation mapping of individual
localized rise in the material temperature followed by the Ti cold sprayed particles deposited at various deposition
material jetting (Ref 4-8). velocity conditions. The variation in mechanical proper-
Much of the current understanding of the cold spray ties within individual cold spray splats and also within
process has been garnered by experimentation and mod- coatings was determined. Nanoindentation technique was
eling of the deposition of Cu (Ref 5, 7-12). Significant used to provide a better understanding of the deposition
interest is placed on the temperature rise near the impact mechanisms of cold spray Ti splats deposited at various
site and especially in the jetting region as this can have a deposition velocity conditions. The effect of the deposi-
profound effect on the final microstructure of cold spray tion velocity conditions on the mechanical properties of
coatings. For Cu, a temperature rise of up to 1200 K was the coatings was investigated.
estimated for the particle impacting at the deposition
velocity of 600 m/s (Ref 4). While the volume of research
on Ti is less, similar studies revealed details of the 2. Experimental Procedure
interface phenomenon for this material (Ref 13-22).
2.1 Sample Preparation
Dina Goldbaum, Richard R. Chromik, and Stephen Yue, Cold Spray deposition was carried out with a Kinetic
Department of Mining and Materials Engineering, McGill 4000 cold spray gun (CGT GmbH, Ampfing, Germany).
University, 3610 University Street, Montreal, QC H3A 2B2,
Nitrogen gas was used as the propelling gas and was
Canada; and Eric Irissou and Jean-Gabriel Legoux, National
Research Council Canada, Industrial Materials Institute, 75 de pressurized from 2 to 4 MPa and preheated to tempera-
Mortagne Blvd, Boucherville, QC J4B 6Y4, Canada. Contact tures ranging from 300 to 800 C. Feedstock powders were
e-mail: richard.chromik@mcgill.ca. introduced into the cold spray gun and passed through a

486—Volume 20(3) March 2011 Journal of Thermal Spray Technology


Peer Reviewed
Fig. 2 Micrograph of the cross-section of a Ti splat obtained by
Fig. 1 Micrograph (SEM) of the CP Ti powder used in this LOM. The flattening ratio was measured from images such as
study these to examine the extent of deformation of Ti splats as a
function of deposition velocity

deLeval nozzle (MOC24, CGT, GmbH, Ampfing,


Germany), reaching in-flight particle velocities between The final polishing step was 0.05 lm colloidal silica. No
580 and 825 m/s. The particle velocity was measured in chemical etching was carried out to minimize surface
free jet at various process conditions with a DPV2000 roughness. The cold spray coatings were prepared in the
(Tecnar Automation, St-Bruno, QC, Canada) in cold same manner, first being cut across the coating cross-
particle mode using a laser diode to illuminate the in-flight section (perpendicular to the gun traverse direction). SEM
particles. The deposition efficiency was measured by tak- imaging was conducted in a variable pressure scanning
ing the ratio of the weight of the deposited coating over electron microscope (Hitachi S-3000 N VP-SEM, Japan).
the weight of the cold sprayed powder. Image analysis of SEM and LOM micrographs of cold
The feedstock powder was commercially pure titanium spray splats and coatings was carried out with commer-
produced by plasma atomization (Raymore Industries, cially available image analysis software (Clemex Vision
Inc., Boisbriand, QC, Canada). The Ti particles had a Professional 5.0, Clemex Technology Inc., Longueuil, QC,
spherical morphology (see Fig. 1) and an average particle Canada). LOM micrographs were used to measure the
diameter of 29 lm. The powder size distribution and splat flattening ratio, FR, by taking a ratio of splat width to
powder microstructure were reported in a previous splat height (Ref 7, 25) as shown in Fig. 2. For each
publication (Ref 22). deposition conditions, the flattening ratio of between five
The cold spray gun was mounted onto an automated and twenty-eight splats was measured. The coating
robotic arm, making possible a wide range of gun motions. porosity was measured from SEM micrographs obtained
Two types of samples were produced by changing the with the instrument in backscattered imaging mode at
gunÕs traverse speed and powder feeding rate. For depo- 2009 magnification. Over three micrographs were ana-
sition of coatings, the powder feeding rate was 20 g/min lyzed and the average of the measurements with a stan-
and the gun traverse speed was kept at 0.33 m/s. Sixteen dard deviation was taken. While porosity measurements
coating layers were deposited on one line, on top of the quoted here are from SEM micrographs, measurements
7.62 9 7.62 9 0.32 cm thick mild steel substrate. Single conducted with LOM micrographs yielded similar results.
splats were deposited on mild steel by changing the gun
traverse speed to 1.0 m/s and by decreasing the powder 2.3 Nanoindentation Testing
feeding rate to 2.0-5.0 g/min. A third type of sample,
multi-pass splats, was produced by increasing the number The nanoindentation testing was carried out with a
of gun passes over the substrate surface. diamond Berkovich tip at a maximum load of 1 mN with
loading and unloading rate of 200 lN/s and a hold time of
1 s. For coatings, profile indentation was carried out with
2.2 Microstructural Characterization
Ubi 3 nanoindentation system (Hysitron Incorporated,
Cold spray splats were examined both in plan view and Minneapolis, MN, USA). A total of 20-40 indents per
in cross-section by scanning electron microscopy (SEM) coating at 20-40 lm indent spacing were performed across
and light optical microscopy (LOM). For cross-section the center of the coating thickness cross-section. For
examination, specimens were mounted in a cold-mount feedstock powders, splats and multi-pass splats, nano-
epoxy. The mounted specimens were prepared with indentation mapping was conducted with a Triboindenter
metallographic techniques recommended for electron (Hysitron Inc., Minneapolis, MN, USA) by indenting a
backscatter diffraction (EBSD), which are designed to matrix of over 100 indents with lateral (x-y) indent spac-
minimize residual stresses induced by polishing damage. ing of 2.5-3.0 lm. Data analysis and graphing software

Journal of Thermal Spray Technology Volume 20(3) March 2011—487


(Origin 8.1, OriginLab Corp., Northampton, MA) was The increase in material deformation and material
Peer Reviewed

used to reconstruct the hardness and reduced modulus jetting with deposition velocity was further revealed by
distribution in terms of a color scale (i.e., maps of hardness examination of splat cross-sections (see Fig. 4a-c). As
and modulus). deposition velocity increases, the flattening ratio increases
The hardness and elastic modulus of the materials were (see Fig. 4d). For a nondeformed powder particle the
calculated from the unloading curves using the Oliver and flattening ratio is 1. The flattening ratio of the splats
Pharr method (Ref 26). The method is based on the deposited over a range of 570 to 825 m/s increases from
analysis of the unloading curve (load versus displacement) 1.8 to 4. The increase from 1 to 1.8 appears to be largely
obtained during sample indentation with a calibrated due to material deformation upon impact (minimal jetting
indenter. For studies presented here, both the machine at lower velocities), while the increase from 1.8 to 4 seems
compliance and the tip area function were determined to be primarily associated with an increase in material
from indentation on a fused quartz standard. jetting. Aside from material jetting other interesting fea-
The material hardness was calculated from Eq. 1, tures were observed that were due to the interaction of Ti
where P is the maximum load applied by the tip and Ac is particles. At a deposition velocity of 636 m/s (see Fig. 5), a
the contact area of the tip. crater formation can be seen at the impact site of the splat.
The impact of the splats resulted in a plastic deformation
P
H¼ ðEq 1Þ in shape of the impacted body, typical of deformation
Ac contributing to a conformal splat adhesion mechanism.
The slope, dP/dh, of the unloading segment defines the While there is also material jetting evident for this splat,
material stiffness, S, as well as the reduced elastic modu- the splat has pulled away and partially de-bonded. At
lus, Er (Ref 26, 27). higher deposition velocities, this de-bonding was not as
pffiffiffiffiffiffi pronounced. In Fig. 6, a splat deposited at 770 m/s also
dP 2Er Ac appears to have partially pulled away. However, in this
S¼ ¼ pffiffiffi ðEq 2Þ
dh p case, a region where the jetting region remains attached is
observed. While impact angle and location of impact has
The contact area, Ac, in Eqs. 1 and 2 was determined from an effect on these observations, with increasing deposition
the calibrated area function of the tip and a contact depth velocity, the number of events resembling Fig. 6 generally
determined from the Sneddon relationship used as part of increased and the number of events resembling Fig. 5
the Oliver and Pharr analysis (Ref 26). decreased. Thus, the increase in jetting with deposition
The reduced elastic modulus consists of an inverse of velocity was a significant mechanism for bonding of par-
the sum of Poisson ratios of the material, tm, and indenter, ticles in multi-pass splats.
ti, divided by their respective elastic modulus, Em and Ei, At higher deposition velocity conditions, the jetting
   1 phenomenon becomes more pronounced and evidence of
1  t2i 1  t2m
Er ¼ þ ðEq 3Þ stronger splat adhesion is observed. Figure 6 shows a Ti
Ei Em splat deposited on a previously deposited Ti splat at the
With the indenter being diamond (Ei = 1120 GPa and deposition velocity of 770 m/s. The close up of the splat
ti = 0.07) and the tested material being Ti (tm = 0.33), one shows a continuous bond between two splats in jetting
finds that the YoungÕs modulus of the tested material, Em region which could point to a formation of the metallur-
is nearly identical to the reduced modulus, Er. For gical bond in that region (Ref 20).
example, if Er = 100 GPa, Em = 102 GPa. The difference
between these two values is much less than the standard 3.2 Microstructure of Cold Sprayed Ti Coatings
deviation on the nanoindentation measurements for Er.
All the values for modulus reported in this article are Figure 7 shows cross-sectional images of three Ti
reduced modulus. Trends or differences in reduced mod- coatings that were produced at deposition velocities of
ulus directly reflect trends or differences in the test 642, 724, and 825 m/s. The coating thickness ranged from
materialÕs YoungÕs modulus. 0.9 to 1.4 mm. The coating porosity decreases with
increase in the deposition velocity and dense coatings with
porosity below 2% are produced at deposition velocity
above 724 m/s. Figure 8(a) shows the effect of the depo-
3. Results and Discussion
sition velocity on the Ti coating porosity. The deposition
efficiency of the coatings increases with increase in the
3.1 Microstructure of Cold Sprayed Splats
deposition velocity and reaches 100% above the deposi-
The appearance of cold spray splats was found to vary tion velocity of 725 m/s. The transition coincides with the
as a function of deposition velocity, revealing features increase in the flattening ratio observed in Fig. 4.
distinctive to the deposition mechanisms. Representative The coatings discussed in this article consist of single
SEM images of splats deposited at velocities between 579 line coatings where the particle impact angle is different
and 825 m/s (see Fig. 3) all showed material jetting, indi- from standard multi-line coatings. Some differences in
cating the presence of adiabatic shear instability (Ref 4). coating porosity measurements and deposition efficiency
However, the jetting phenomenon increased with increase are expected between coatings produced here and those
in the splat deposition velocity. produced at similar conditions, but as multi-line coatings

488—Volume 20(3) March 2011 Journal of Thermal Spray Technology


Peer Reviewed
Fig. 3 Micrographs (SEM) of CP Ti splats deposited at: (a) 300 C, 2 MPa (580 m/s), (b) 300 C, 4 MPa (642 m/s), (c) 500 C, 2 MPa
(636 m/s), (d) 500 C, 4 MPa (724 m/s), (e) 750 C, 3 MPa (770 m/s), and (f) 800 C, 4 MPa (825 m/s) gas preheat temperature, gas
pressure and particle velocity

(Ref 22). Single line coatings were chosen for this study hardness and modulus color scale are annotated by gray
because the particle jet center was easily connected to the and black, while white indicates missing data points. In
center of the coating cross-section, avoiding any concerns Fig. 9, black regions represent the measurements made in
over variation of velocity within the jet affecting our the epoxy which has low hardness, below 2.2 GPa, and
measurements (Ref 2). modulus, below 80 GPa. The regions in gray can be found
in the mild steel substrate that has a higher elastic mod-
ulus when compared to titanium. For comparison, the
3.3 Nanoindentation Mapping of Cold Sprayed
optical images of the indented cold spray splats are shown
Splats and Multi-Pass Splats
at the left-hand side of Fig. 9.
Mechanical property mapping was carried out on the Nanoindentation maps reveal that the hardness distri-
cold spray Ti splats and feedstock Ti powder. Figure 9 bution inside of the cold spray splats or feedstock powder
depicts the distribution of hardness and modulus of feed- is not homogeneous (see Fig. 9). In the feedstock powder,
stock Ti powder and cold spray splats deposited at 642, high hardness of 3.6 to 3.8 GPa is measured at the
724, and 825 m/s. The color notation is interpolated powder outer surface layer. A martensitic microstructure
between the measurements from indented data points. was often observed in this region, possibly a result of
The color scale has a gradient of 0.2 GPa for hardness and surface tension during the powder production through
10 GPa for modulus. The upper and lower limits of the a plasma atomization process. However, in splats,

Journal of Thermal Spray Technology Volume 20(3) March 2011—489


Peer Reviewed

Fig. 4 Micrographs (LOM) of Ti splat cross-section deposited at (a) 300 C, 4 MPa (642 m/s), (b) 500 C, 4 MPa (724 m/s), (c) 800 C,
4 MPa (825 m/s) gas preheat temperature, gas pressure, and particle velocity, and (d) the flattening ratio with standard deviation of Ti
splats with respect to the deposition velocity

Fig. 5 Micrographs (SEM) of splat deposited at 500 C, 2 MPa (636 m/s) and demonstrating a conformal particle adhesion

high hardness regions are mainly observed at the splat expected to cause an increase in hardness but have little
impact site with the substrate material. The hardness in effect on the elastic properties.
these regions reaches 4.2 GPa and is color coded in red. As opposed to the single pass splats, multi-pass splats
Relatively lower hardness of 3.0-3.4 GPa is measured in showed a higher modulus. In Fig. 10, modulus of 120 GPa
the upper splat portion. and higher is measured in the splats deposited above
In Table 1, the average hardness and modulus of the 500 C gas preheat temperature. As the gas preheat tem-
splats have been calculated. The average splat hardness perature increases to 800 C, the modulus of the splats
does not increase significantly with increase in the depo- increases up to 130 GPa. These values of the modulus are
sition velocity. The splats hardness is, however, higher consistent with that of Ti, and could simply be an effect of
than that of the feedstock powder. The modulus of the differences in crystallographic texture.
single splats is similar to the modulus of the feedstock In multi-pass splats, higher hardness is found both at
powder. The residual stresses induced by impact are the impact site between splats and near the steel substrate.

490—Volume 20(3) March 2011 Journal of Thermal Spray Technology


Peer Reviewed
Fig. 6 Micrographs (SEM) showing ductile fracture in the jetting region of CP Ti splat deposited at 750 C, 3 MPa (770 m/s)

Fig. 7 Micrographs (SEM) of cold spray Ti coatings deposited at: (a) 300 C, 4 MPa (642 m/s), (b) 500 C, 4 MPa (724 m/s), and
(c) 800 C, 4 MPa (825 m/s)

In Fig. 10, multi-pass titanium splats deposited at 642, 724, spray coatings did not change with the coating thickness.
and 825 m/s are shown along with hardness and modulus A typical profile of hardness and modulus measured
maps. Regions of 3.7-4.2 GPa hardness were measured in across coating thickness cross-section is shown in
the impact region between two splats as well as in the Fig. 11(a) for a coating deposited at 500 C gas preheat
impact region of the splat on the steel substrate. However, temperature, 4 MPa gas pressure, and 724 m/s particle
in the jetting region, a lower hardness of 3.2-3.4 GPa was deposition velocity. Similar trends were observed for all
measured for the 825 m/s splat. coatings produced at all deposition velocity conditions.
This result is contrary to previously published results for
microindentation on cold sprayed titanium (Ref 11, 22)
3.4 Nanoindentation of Cold Sprayed Ti Coatings
where the hardness was seen to increase near the sub-
Profile nanoindentation across cold spray Ti coatings strate/coating interface due to a tamping effect. This
revealed that the mechanical properties of these cold difference is likely due to the effect porosity and particle

Journal of Thermal Spray Technology Volume 20(3) March 2011—491


Peer Reviewed

Fig. 8 Graphs of (a) porosity with a standard deviation and (b) deposition efficiency of cold spray Ti coatings vs. deposition velocity

Fig. 9 Hardness and modulus maps of Ti powder for Ti splats deposited at: 300 C, 3 MPa (642 m/s); at 500 C, 4 MPa (724 m/s), and at
800 C, 4 MPa (825 m/s). Micrographs (LOM) of the splats and powder are provided on the left-hand side. The hardness and modulus color
scale bars are provided at the right-hand side with black demonstrating lower end values and gray showing higher end values

de-bonding has on the hardness test. Nanoindentation is Ti powder plotted at 0 m/s. Higher hardness and modulus
affected the least and represents mechanical properties of are measured in the cold spray coatings in comparison to
the cold sprayed material, while microindentation pro- the titanium feedstock powder. The hardness and modulus
vides a measure of the overall mechanical response of the of the coatings deposited above 625 m/s was 3.7 ± 0.2
coating, including defects. GPa and 125 ± 10 GPa, respectively. The hardness and
The average hardness and modulus from profile modulus of the initial feedstock powder was 3.2 ± 0.2 GPa
indentation is plotted against the deposition velocity in and 110 ± 5 GPa. No significant change in the coating
Fig. 11(b), with the mechanical properties of the feedstock hardness with increase in the coating deposition velocity

492—Volume 20(3) March 2011 Journal of Thermal Spray Technology


Table 1 Mechanical properties of Ti powder and cold spray Ti coatings and Ti splats

Peer Reviewed
Deposition Deposition Coating Coating Coating Splat Splat Splat
velocity, efficiency, porosity, hardness, modulus, flattening hardness, modulus,
Specimen/conditions m/s % % GPa GPa ratio GPa GPa

Ti powder/as-polished N/a N/a <1 3.1 ± 0.3 107 ± 9 N/a 3.1 ± 0.3 107 ± 9
Ti/300 C, 2 MPa 580 19 18.0 ± 10.1 3.2 ± 0.3 99 ± 11 1.8 ± 0.1 - -
Ti/300 C, 3 MPa 625 30 2.5 ± 0.3 3.6 ± 0.2 112 ± 12 2.0 ± 0.2 3.4 ± 0.4 103 ± 9
Ti/300 C, 4 MPa 642 49 5.9 ± 1.5 3.7 ± 0.2 116 ± 13 2.2 ± 0.3 - -
Ti/500 C, 2 MPa 636 76 7.4 ± 1.6 3.6 ± 0.4 117 ± 10 2.4 ± 0.5 - -
Ti/500 C, 3 MPa 694 83 1.4 ± 0.6 3.7 ± 0.3 120 ± 11 2.5 ± 0.5 - -
Ti/500 C, 4 MPa 724 91 2.1 ± 0.6 3.5 ± 0.3 109 ± 8 2.3 ± 0.4 3.6 ± 0.4 119 ± 6
Ti/750 C, 3 MPa 770 100 0.7 ± 0.1 3.7 ± 0.4 120 ± 7 3.0 ± 0.6 - -
Ti/800 C, 4 MPa 825 100 1.3 ± 0.3 3.5 ± 0.3 121 ± 6 3.9 ± 1.0 3.7 ± 0.4 110 ± 7

Fig. 10 Hardness and reduced modulus maps of the multi-pass splats deposited at 300 C, 3 MPa (642 m/s); at 500 C, 4 MPa (724 m/s),
and at 800 C, 4 MPa (825 m/s). Micrographs (LOM) of the splats and powder are provided on the left-hand side. The hardness and
modulus color scale bars are provided at the right-hand side with black demonstrating lower end values and gray showing higher end
values

past 650 m/s range was observed. Average hardness 3.5 Deposition Mechanisms, Mechanical
and modulus for all the coatings are summarized in Properties, and Comparisons to Literature
Table 1.
Splat bonding mechanisms, extent of deformation,
It should be noted that mechanical property mapping
degree of jetting, and mechanical properties were all
was also conducted on coatings. Similar small-scale vari-
observed to vary to greater or lesser degrees as a function
ation in mechanical properties across particle boundaries
of deposition velocity. The relationships between all these
was also observed. Thus, while the average properties of
phenomena are complex, but the results of this study
the cold sprayed coatings do not change over the length
along with those in the literature leads to the following
scale of the coating thickness, small variations similar to
discussion of the observable trends with deposition
those observed in the splats do exist in the coatings at a
velocity.
similar length scale to the particles.

Journal of Thermal Spray Technology Volume 20(3) March 2011—493


Peer Reviewed

Fig. 11 (a) Nanoindentation profile hardness and modulus of Ti coatings deposited at 500 C, 4 MPa (724 m/s) away from the coating
and the substrate interface plotted at 0 mm, (b) Average profile nanoindentation hardness and modulus with standard deviation of
coatings plotted against their respective deposition velocities. For reference, the average properties for the powder are plotted at 0 m/s

When titanium was sprayed at velocities below 725 m/s,


the deposition efficiency was below 100% and the flat-
tening ratio of the splats was roughly constant between 1.8
and 2.5. While some material jetting was observed for
these splats, it was somewhat minimal compared to those
deposited at higher velocities (c.f. Fig. 3, 9). Additionally,
evidence of de-bonding of splats (c.f. Fig. 5) indicated that
the jetting was in some instances not providing a strong
metallurgical bond. Above 725 m/s, the deposition effi-
ciency was 100% and the flattening ratio rose to a value of
3 at 770 m/s and 3.9 at 825 m/s. In this regime, significant
jetting occurred and evidence metallurgical bonding was
also found (c.f. Fig. 6). These observations, especially the
increase in material jetting with particle velocity, are
consistent with the current understanding of the mecha-
nism of adiabatic shear instability as a deposition mecha- Fig. 12 Hardness distribution with respect to the splat deposi-
tion velocity in Ti splat regions: near interface, in the jetting
nism for cold spray (Ref 4, 6, 7). The transition to 100% region and in the upper splat region
deposition efficiency at roughly 725 m/s matches very well
the predicted critical velocity of ~700 m/s for 25 lm
diameter titanium particles, Schmidt et al. (Ref 4). In also evident in the variation of hardness in the different
terms of mechanical properties, the cold spray splats and regions of the splats. This discussion and the splat hardness
coatings are always on average harder than the feedstock maps correlate well with the dislocation density distribu-
Ti powder. However, no significant increase in the coating tion in the TEM image of Ti splats taken by Kim et al.
or splat hardness was measured with the increase in par- (Ref 24) and also the simulated stress distribution patterns
ticle velocity. To explore this phenomenon further, we proposed by Assadi et al. (Ref 7) Thus, there appears to be
plot the hardness of cold spray splats near the interface a consistent picture of high dislocation density and hard-
and far from the interface as a function of deposition ness near the interface, minimal activity in the upper splat
velocity (see Fig. 12). Near interface hardness is generally region, and the possibility of thermal softening in the
significantly higher than the feedstock powder, indicating jetting regions.
work hardening during the impact. This hardness remains According to Kapoor et al., high strain rate deforma-
relatively constant with particle velocity, seeming to tion of Ti, such as that occurring in the jetting region, can
indicate that the extent of work hardening has reached a be very efficient at converting deformation work into heat
maximum in this area even at the lowest velocity. How- (Ref 25). The temperature rise in the jetting region was
ever, as the velocity increases, the hardness far from the estimated to reach near melting point (<90% of Tm) for Ti
interface does increase gradually, perhaps due to some splats deposited on a steel substrate with nitrogen gas at
deformation and increased dislocation density in the 950 m/s deposition velocity (Ref 20). Higher temperatures
upper splat regions. Lastly, the hardness of the jetting have been estimated for Ti splats deposited on a Ti sub-
region is also plotted in Fig. 12 for the splat with the strate (Ref 28). The extent to which this heat affects the
highest deposition velocity of 825 m/s (c.f. Fig. 10). In the resulting microstructure and distribution of strain is diffi-
jetting region, the hardness is similar to the feedstock cult to predict. Others have shown that the temperature
powder. Thus, while the average mechanical properties rise during impact of cold sprayed material is sufficient to
remain constant with deposition velocity, the complex induce a dynamic recrystallization in the cold spray
phenomenon of particle impact, deformation and jetting is deposited Ti splats (Ref 24). In addition, there has been

494—Volume 20(3) March 2011 Journal of Thermal Spray Technology


evidence that the recrystallization can result in preferen- technical assistance of Ahmad Rezaeian, Bernard Harvey

Peer Reviewed
tial grain orientation of cold spray Ti coatings (Ref 14). and Frederic Belval.
For all splats and coatings studied here, the modulus
measured by indentation is within the normal range for Ti,
which can vary from 100 to 145 GPa (c.f. Eq. 3). As it is References
not a pure uniaxial test, indentation is not very sensitive to
changes variations due to crystallographic orientation 1. E. Irissou, J.G. Legoux, A. Ryabinin, B. Jodoin, and C. Moreau,
Review on Cold Spray Process and Technology: Part I, Intellec-
(Ref 29, 30). Thus, the small changes observed in our data tual Property, J. Therm. Spray Technol., 2008, 17(4), p 495-516
may or may not reflect some degree of preferred orien- 2. A. Papyrin, V. Kosarev, K. V. Klinkov, A. Alkhimov, and
tation in the sprayed material. V. M. Fomin, Cold Spray Technology, Elsevier, Oxford, 2006,
p 74, 153-169, 259
3. V.K. Champagne, The Cold Spray Materials Deposition Process:
Fundamentals and Applications, Woodhead Publishing Limited,
Cambridge, 2007, p 362
4. Conclusions 4. T. Schmidt, F. Gärtner, H. Assadi, and H. Kreye, Development
of a Generalized Parameter Window for Cold Spray Deposition,
The mechanical properties of cold sprayed Ti splats and Acta Mater., 2006, 54(3), p 729-742
5. C.-J. Li, W.-Y. Li, and H. Liao, Examination of the Critical
coatings were studied by nanoindentation. When com- Velocity for Deposition of Particles in Cold Spraying, J. Therm.
pared to the hardness of the feedstock powder, the dis- Spray Technol., 2006, 15(2), p 212-222
tribution of hardness within individual Ti cold spray splats 6. M. Grujicic, C.L. Zhao, W.S. DeRosset, and D. Helfritch, Adi-
demonstrated three regions: (1) an impact site with an abatic Shear Instability Based Mechanism for Particles/Substrate
increase in hardness, (2) an upper splat region with similar Bonding in the Cold-Gas Dynamic-Spray Process, Mater. Des.,
2004, 25(8), p 681-688
hardness to feedstock material, and (3) jetting regions 7. H. Assadi, F. Gärtner, T. Stoltenhoff, and H. Kreye, Bonding
where the hardness is similar to the feedstock material. Mechanism in Cold Gas Spraying, Acta Mater., 2003, 51(15),
At lower deposition velocities (<650 m/s), the first two p 4379-4394
regions were more evident than the third, indicating pri- 8. T. Schmidt, F. Gaertner, and H. Kreye, New Developments in
Cold Spray Based on Higher Gas and Particle Temperatures,
marily conformal adhesion. At higher deposition veloci- J. Therm. Spray Technol., 2006, 15(4), p 488-494
ties (>650 m/s), the third region, the jetting region, 9. C. Borchers, F. Gartner, T. Stoltenhoff, and H. Kreye, Micro-
became more pronounced, indicating the formation adia- structural Bonding Features of Cold Sprayed Face Centered
batic shear instability. The mechanical properties mea- Cubic Metals, J. Appl. Phys., 2004, 96(8), p 4288-4292
sured in these three regions are consistent with results in 10. T. Kairet, M. Degrez, F. Campana, and J.P. Janssen, Influence of
the Powder Size Distribution on the Microstructure of Cold-
the literature that show increased dislocation density near Sprayed Copper Coatings Studied by X-ray Diffraction, J. Therm.
the interface and dynamic recrystallization in the jetting Spray Technol., 2007, 16(5), p 610-618
region. Thus, at the highest deposition velocities, which 11. A. Rezaeian, E. Irissou, R.R. Chromik, S. Yue, Characterization
are preferred due to low coating porosity and high depo- of Cold-Sprayed Ni, Ti, Cu Coating Properties for their
Optimization, Thermal Spray: Crossing Borders, Jun 2-4, 2008
sition efficiency, there was sufficient heat in the jetting (Maastricht, The Netherlands), ASM International, p 854-859
regions to promote dynamic recrystallization in Ti and a 12. T.H. Van Steenkiste, J.R. Smith, R.E. Teets, J.J. Moleski,
reduction in hardness. However, there is retention of work D.W. Gorkiewicz, R.P. Tison, D.R. Marantz, K.A. Kowalsky,
hardening at the impact site away from the jetting region, W.L. Riggs, P.H. Zajchowski, B. Pilsner, R.C. McCune, and
where there is high strain, but insufficient strain rate to K.J. Barnett, Kinetic Spray Coatings, Surf. Coat. Technol., 1999,
111(1), p 62-71
realize a significant temperature rise. Despite the signifi- 13. T. Stoltenhoff, C. Borchers, F. Gärtner, and H. Kreye, Micro-
cant differences in hardness in the different regions, the structures and Key Properties of Cold-Sprayed and Thermally
average hardness of cold sprayed Ti splats remained Sprayed Copper Coatings, Surf. Coat. Technol., 2006, 200(16-17),
similar for all deposition velocities and was greater than p 4947-4960
14. S. Zahiri, D. Fraser, and M. Jahedi, Recrystallization of Cold
the feedstock material. The same was observed for the Spray-Fabricated CP Titanium Structures, J. Therm. Spray
cold spray coatings produced, where there were local Technol., 2009, 18(1), p 16-22
variations in mechanical properties but the coating 15. C.-J. Li and W.-Y. Li, Deposition Characteristics of Titanium
mechanical properties remained constant across the Coating in Cold Spraying, Surf. Coat. Technol., 2003, 167(2-3),
coating thickness cross-section. p 278-283
16. W.Y. Li, C. Zhang, X. Guo, J. Xu, C.J. Li, H. Liao, C. Coddet,
and K.A. Khor, Ti and Ti-6Al-4V Coatings by Cold Spraying and
Microstructure Modification by Heat Treatment, Adv. Eng.
Mater., 2007, 9(5), p 418-423
Acknowledgments 17. R.S. Lima, A. Kucuk, C.C. Berndt, J. Karthikeyan, C.M. Kay, and
J. Lindemann, Deposition Efficiency, Mechanical Properties and
Financial support from the Canadian Foundation for Coating Roughness in Cold-Sprayed Titanium, J. Mater. Sci.
Innovation is gratefully acknowledged; the cold spray Lett., 2002, 21(21), p 1687-1689
equipment was provided by CFI project No. 8246 while 18. T. Marrocco, D. McCartney, P. Shipway, and A. Sturgeon, Pro-
the nanoindentation equipment was provided by CFI, duction of Titanium Deposits by Cold-Gas Dynamic Spray:
LeaderÕs Opportunity Fund, project No. 13029. Opera- Numerical Modeling and Experimental Characterization,
J. Therm. Spray Technol., 2006, 15(2), p 263-272
tional funding for this project was provided by the Natural 19. J. Vlcek, H. Huber, H. F. Voggenreiter, and E. Lugscheider,
Sciences and Engineering Research Council (NSERC) Melting Upon Particle Impact in the Cold Spray Process, Inter-
Strategic Grants Program. The authors acknowledge the national Congress on Advanced Materials, Their Process and

Journal of Thermal Spray Technology Volume 20(3) March 2011—495


Applications, September 30-October 2, 2002 (Munich, Germany), 25. R. Kapoor and S. Nemat-Nasser, Determination of Temperature
Peer Reviewed

Deutsche Gesellschaff Fur Materialkunde (DSM) Rise During High Strain Rate Deformation, Mech. Mater., 1998,
20. G. Bae, S. Kumar, S. Yoon, K. Kang, H. Na, H.J. Kim, and C. Lee, 27(1), p 1-12
Bonding Features and Associated Mechanisms in Kinetic Sprayed 26. A. C. Fisher-Cripps, Nanoindentation, 2nd ed., Springer,
Titanium Coatings, Acta Mater., 2009, 57(19), p 5654-5666 New York, 2002, p 1-20
21. K. Kim, M. Watanabe, K. Mitsuishi, K. Iakoubovskii, and S. Kuroda, 27. W.C. Oliver and G.M. Pharr, Measurement of Hardness and
Impact Bonding and Rebounding Between Kinetically Sprayed Elastic Modulus by Instrumented Indentation: Advances in
Titanium Particle and Steel Substrate Revealed by High-Resolution Understanding and Refinements to Methodology, Mater. Res.
Electron Microscopy, J. Appl. Phys. D, 2009, 42(6), p 65304 Soc., 2004, 19(1), p 3-20
22. W. Wong, A. Rezaeian, E. Irissou, J.G. Legoux, and S. Yue, Cold 28. G. Bae, Y. Xiong, S. Kumar, K. Kang, and C. Lee, General
Spray Characteristics of Commercially Pure Ti and Ti-6Al-4V, Aspects of Interface Bonding in Kinetic Sprayed Coatings, Acta
Adv. Mater. Res., 2010, 89-91, p 639-644 Mater., 2008, 56(17), p 4858-4868
23. K.-H. Kim, M. Watanabe, J. Kawakita, and S. Kuroda, Grain 29. J.J. Vlassak and W.D. Nix, Measuring the Elastic Properties of
Refinement in a Single Titanium Powder Particle Impacted at Anisotropic Materials by Means of Indentation Experiments,
High Velocity, Scripta Mater., 2008, 59(7), p 768-771 J. Mech. Phys. Solids, 1994, 42(8), p 1223-1245
24. K.-H. Kim, M. Watanabe, J. Kawakita, and S. Kuroda, Effects of 30. J.J. Vlassak, M. Ciavarella, J.R. Barber, and X. Wang, The
Temperature of In-flight Particles on Bonding and Microstruc- Indentation Modulus of Elastically Anisotropic Materials for
ture in Warm-Sprayed Titanium Deposits, J. Therm. Spray Indenters of Arbitrary Shape, J. Mech. Phys. Solids, 2003, 51(9),
Technol., 2009, 18(3), p 392-400 p 1701-1721

496—Volume 20(3) March 2011 Journal of Thermal Spray Technology

You might also like